A Case Study on Atlantic Tropical Cyclogenesis and Saharan Air... Using WRF/Chem Coupled with an AOD Data Assimilation System

advertisement
A Case Study on Atlantic Tropical Cyclogenesis and Saharan Air Layer Simulated
Using WRF/Chem Coupled with an AOD Data Assimilation System
Zhiquan Liu and Dan Chen
National Center for Atmospheric Research, Boulder, Colorado, USA
Abstract
This study investigated the dust radiative effects on tropical cyclogenesis and the Saharan Air
Layer (SAL) over the Atlantic Ocean by using the Gridpoint Statistical Interpolation threedimensional variational data assimilation (DA) system coupled with the Weather Research
and Forecasting/Chemistry model. Two experiments were conducted with (DA) and without
(NO_DA) the assimilation of MODIS 550-nm aerosol optical depth (AOD) retrievals. Oneweek (from 28 August to 5 September, 2006) assimilation of MODIS AOD substantially
improved the depiction of Saharan dust outbreak and transport feature. Differences of 180-hr
forecasts initiated at 00 UTC 5 September from the two experiments were analyzed to
evaluate the dust radiative effects. The dust-radiation interaction over mostly cloudless
condition in the first 3-day model simulation mainly warmed the dusty layer and the layer
near the surface via dominated dust absorption effect of shortwave solar radiation. At a later
stage of forecasts (Sep. 9-11) with cloudier conditions and more pronounced semi-direct
impact, the Meso-sclae Convective Systems (MCS) have developed over time in the Main
Cyclogenesis Region with significant temperature increase and relative humidity decrease in
the middle troposphere when assimilating AOD. This suppressed clouds and precipitation and
enhanced downward motion and atmospheric stability, which finally resulted in the erosion of
spurious MCSs appeared in the NO_DA experiment. Due to early suppression, more ice
precipitation and high clouds were formed as more water ascended to freezing levels, and the
greater amount of freezing water aloft released extra latent heat and invigorated the
convection in peripheral region. The storm was further weakened by the bled and cooled lowlevel airflow caused by peripheral clouds invigoration. Near the end of 180-hr forecast from
the AOD DA experiment, the hotness and dryness features of the SAL were more pronounced
when compared to the NO_DA experiment and likely reflected a more realistic model
simulation of dust-radiative effects.
1 1. Introduction
Mineral dust, one of the most abundant aerosol species in the atmosphere, has important
weather and climatic effects through its influence on solar and terrestrial radiation and the
radiative and physical properties of clouds [e.g., Sokolik et al., 1999; Ginoux et al., 2001;
Ramanathan et al., 2001; Lau et al., 2009; Zhao et al., 2010]. The Sahara desert over North
Africa is the largest source of mineral dust in the world. During the summer months, Saharan
dust outbreaks are associated with a dry and hot well-mixed layer (Saharan Air Layer, SAL)
extending from 1500 m to 6000 m over the North Atlantic Ocean. Several studies have
examined the Saharan dust radiative effects on the SAL and environment shear [e.g., Chen et
al., 2010; Zhao et al., 2010]. For dust plumes under cloudless conditions over the eastern
Atlantic Ocean, the radiative effect is dominated by the dust shortwave radiation interaction,
resulting in net heating with the maximum occurring slightly below the peak dust
concentration. High aerosol optical depth (AOD) below 700 hPa is associated with increased
atmospheric stability through increased temperature and dryness. Thus in addition to the hot
air that originates from the desert upstream, the dust-induced radiative forcing helps the SAL
to retain the warmness and dryness throughout its depth as it is carried as far as the western
Caribbean Sea.
Saharan dust, often propagating downstream along SAL to the Atlantic Ocean, can modify the
SAL and its environment by changing the energy budget through direct, semi-direct and
indirect effects [e.g., Chen et al., 2010; Su et al., 2008; Zhao et al., 2010]. Tropical cyclones
(TCs) in the Atlantic basin often develop from meoscale convective systems (MCSs)
embedded within African easterly waves (AEWs) that originate over West Africa [Landsea,
1993]. Near the southern boundary of SAL, the dusty, hot and dry air is associated with
AEWs, which can affect TC activity. In cloudy conditions where relatively cool and moist
marine air undercuts the SAL as it moves westward, the dust can affect clouds through semidirect thermodynamic process or indirect cloud microphysical processes. Several
observational and numerical studies [e.g., Karyampudi and Carlson, 1988; Karyampudi and
Pierce, 2002; Dunion and Velden, 2005; Evan et al., 2006; Jones et al., 2007; Wu, 2007; Shu
and Wu, 2009; Sun et al., 2008, 2009; Partt and Evans, 2009; Reale et al, 2009, Chen et al.,
2010] have shown evidence for aerosol-induced intensification and weakening of TC
development. For example, by using satellite data Dunion and Velden [2005] found that the
dry SAL can suppress Atlantic TC activity by increasing the vertical wind shear and
stabilizing the environment at low atmospheric levels. They also suggested that convectively
driven downdrafts caused by the SAL dry air can be an important inhibiting factor for TCs.
To accurately forecast TC intensity, research activities and field experiments have also
examined tropical cyclogenesis [Zipser et al., 2009]. Tropical cyclogenesis is strongly
influenced by environmental factors, such as the sea surface temperature, moisture in the
lower troposphere and vertical wind shear [Gray, 1968]. Sun et al. [2009] and Wu et al. [2007]
incorporated the Atmospheric Infrared Sounder (AIRS) measured temperature and humidity
2 profiles into numerical weather prediction models (WRF and MM5 respectively), to study the
effects of SAL on TC activity. Both of them found that with the assimilation of AIRS
observations during dust outbreak periods, the dry and warm SAL was better simulated such
that the development of tropical disturbances was inhibited along the southern edge of the
SAL. In their studies, the effects on TC development were investigated from the view of SAL.
The magnitude of the radiative impact of dust highly depends upon the distribution of the dust
concentration and its optical characteristics. While the modeling of the spatial distribution of
Saharan dust and its optical depth remains uncertain and challenging, AOD data assimilation
(DA), combining satellite derived AOD observations with numerical model output, has
proved to be skillful at improving aerosol and AOD forecasts [Collins et al., 2001; Liu et al.,
2011]. Liu et al. [2011] implemented AOD DA within the National Centers for Environmental
Prediction (NCEP) Gridpoint Statistical Interpolation (GSI) three-dimensional variational
(3DVAR) DA system coupled to the Goddard Chemistry Aerosol Radiation and Transport
(GOCART) [Chin et al., 2000, 2002] aerosol scheme within the Weather Research and
Forecasting/Chemistry (WRF/Chem) model [Grell et al., 2005]. Their results demonstrated
improved aerosol forecasts from AOD DA over a week-long period while studying a dust
storm event over East Asia.
In this study, we will investigate the dust-radiation effects on SAL and TC genesis in the
Atlantic Ocean using the GSI-WRF/Chem coupled AOD DA system. The Atlantic TC case
and motivation are presented in section 2, followed by model description and experimental
design in section 3. Model simulation results are described in section 4 and conclusions are
given in section 5.
2. Case and Motivation
We selected the same period as in Sun et al. [2009], which spanned Sep. 5-12, 2006. AIRS
retrievals of vertical temperature (T) and relative humidity (RH) profiles showed the clear
feature of SAL - 1-4 K higher T and 50% lower RH in the low-middle atmosphere (e.g., 850700 hPa) of the heavy dusty area (dust AOD > 0.4).
It is likely that higher T observed by AIRS in the heavy dust area was induced by dust
radiative effects. In this study, we attempt to investigate three questions: 1. Can the feature of
the Saharan dust outbreak be well simulated by assimilating MODIS AOD observations? 2.
How does dust affect TC development through radiative effects? Can the TC suppression be
reproduced as in Sun et al [2009]? 3. Can the dry and warm SAL be better simulated with the
assimilated dust field? To the authors’ knowledge, this is the first study to investigate the TC
genesis and SAL using the WRF/Chem model.
3. Model Description and Experimental Design
3 Version 3.4.1 of WRF/Chem was used with “online” coupled meteorological and chemical
processes. The model domain with 36-km horizontal grid spacing covers the Atlantic Ocean
and Sahara desert (Fig. 1). There are 57 vertical levels extending from the surface to 10 hPa.
Aerosol direct and semi-direct effects are allowed in WRF/Chem by linking the optical
properties of GOCART aerosols (OC and BC with 2 size bins, sulfate, dust with 5 size bins
and sea salt with 4 size bins) to the Goddard Space Flight Center Shortwave radiation scheme
[Chou and Suarez, 1994]. Aerosol indirect effects were not implemented for GOCART with
the WRF/Chem version used. The WRF single-moment 6-class microphysics scheme and the
Grell-Devenyi ensemble cumulus parameterization [Grell and Devenyi, 2002] were used. The
dust emission flux is computed as a function of probability source function and surface wind
speed [Ginoux et al., 2001]. Similar to dust uplifting, sea salt emissions from the ocean are
highly dependent on the surface wind speed and calculated as a function of wind speed at 10
m and sea salt particle radius [Chin et al., 2002]. The lateral boundary conditions (LBCs) for
meteorological fields were provided by the NCEP GFS forecasts. LBCs for chemistry/aerosol
fields were idealized profiles embedded within the WRF/Chem model and represented clean
oceanic conditions (could cite Schwartz et al. [2012] if you wanted).
NCEP’s GSI 3DVAR DA system was used to assimilate total 550-nm AOD (hereafter “AOD”)
retrievals from MODIS sensors onboard Terra and Aqua satellites, as described in Liu et al.
[2011]. MODIS AOD retrievals of Deep Blue products over desert [Hsu et al., 2004] and
dark targeting products over ocean and vegetated land [Remer et al., 2005] were assimilated in
this study. GSI 3DVAR system calculates a best-fit “analysis” considering the observations
(AOD in our case) and background fields (short-term WRF/Chem forecasts in our case)
weighted by their error characteristics. The community radiative transfer model (CRTM) [Han
et al., 2006; Liu and Weng, 2006] is used as the AOD observation operator in GSI to
transform the GOCART aerosol profiles into AOD. Each 3-D aerosol species is used as
analysis variables of 3DVAR, and the 3-D mass concentrations of the aerosol species are
analyzed in a one-step minimization procedure constrained by the observation and
background error covariances (BECs). The AOD observation error is assumed to be spatially
uncorrelated and modeled as in Liu et al. [2011]. To more accurately reflect the dustdominated feature, the BEC statistics for each aerosol variable were calculated by utilizing the
“NMC method” [Parrish and Derber, 1992] based upon one-month of WRF/Chem forecasts
for the dust outbreak month of August 2005. Standard deviations and horizontal/vertical
correlation lengthscales of the background errors (separated for each aerosol species) were
calculated using the method described by Wu et al. [2002]. Further details regarding the AOD
DA system, including the algorithm, observation operator, and modeling of the background
and observation error covariances can be found in Liu et al. [2011].
Two experiments were conducted using this coupled GSI-WRF/Chem system with (DA) and
without (NO_DA) assimilating MODIS AOD observations. Both experiments initialized a 6hr WRF/Chem forecast every 6-hr starting from 00 UTC 28 August 2006. Meteorological
4 initial conditions (ICs) for each forecast came from GFS analyses. For NO_DA, aerosol fields
were simply carried over from cycle to cycle (similar to a continuous aerosol forecast). For
DA, GSI 3DVAR updated GOCART aerosol variables by assimilating MODIS AOD at 12
UTC and 18 UTC (when AOD observations were available) using the previous cycle’s 6-hr
forecast as the background. A final analysis was made at 00 UTC 5 September from this 6
hourly cycling procedure and then a 180-hr forecast was initialized for both experiments. Note
that meteorological ICs in the two 180-hr forecasts are the same (from the GFS analysis) and
the forecast differences of meteorological fields are resulted solely from the differences of
aerosol ICs between the DA and NO_DA experiments. We now present results from the two
180-hr forecasts.
4. Results
4.1 Assimilated Dust Outbreak
After a one-week period (00 UTC 28 August – 00 UTC 5 September) for aerosol spin-up,
the impact of AOD DA on the depiction of the dust outbreak event can be clearly seen in
Figure 1, which shows the spatial distribution of AOD valid at 00 UTC 5 September from the
NO_DA and DA experiments. The ratio of dust contributed AOD is also shown as red contour
lines. While the NO_DA experiment substantially underpredicted aerosol concentrations,
particularly over the Atlantic Ocean, the DA experiment produced much larger AOD values
than the NO_DA experiment not only in the dust source region but also over the Atlantic
Ocean (~1.3 vs. ~0.25 for the maximum values). In the DA experiment, the high AOD values
extended westward to 60° W and spanned about 10-30 degrees in latitude off the coast of
northern Africa. While the dust AOD ratios over ocean reached as high as 0.8 in the DA
experiment, clearly indicating a dust-dominated event, the NO_DA experiment had the dust
AOD ratios less than 0.4 over ocean, significantly underpredicting the dust outbreak. The
feature in the DA experiment is consistent with the OMI Aerosol Index images shown in Sun
et al. [2009].
Note that only column-total 550-nm AOD was assimilated and no aerosol speciation
information was contained in observations. It is important to have the phenomena-specific
background error statistics to allow an appropriate adjustment of individual aerosol species
during assimilation. These error statistics were obtained using the WRF/Chem forecasts for
the dust outbreak month of August 2005. Consistent with the dust outbreak period, the
standard deviations of dust errors are one- or two-orders of magnitude larger than those of
other species (not shown). Larger background errors of dust allowed larger adjustments of the
dust field, which is crucial for the aerosol analyses of the Sahara dust outbreak period in this
study.
4.2 Impact on Tropical Cyclogenesis
5 As our main goal is to investigate the dust-radiative effects on TC development in the 180-h
forecasts initialized from 00 UTC 5 September, a small region (black rectangle in Fig. 1b)
defined as Main Cyclogenesis Region (MCR), where TCs often develop as the dust resided, is
ideal for a detailed analysis of results. As TCs in the Atlantic basin often develop from MCSs
embedded within AEWs, cloud water and convective precipitation are good indicators of
MCSs. The Eastern Atlantic Ocean was mostly under cloudless conditions in the first 48-hr
forecasts (not shown). The convective system developed over time and obvious clouds
occurred after 8 September.
Figure 2 shows the MCR-averaged differences of AOD, vertical velocity, T, RH, cloud liquid
and cloud ice water as a function of height and forecast valid time between the two
experiments (DA – NO_DA). The time series started from 00 UTC Sep 5. The largest AOD
increase due to dust outbreak occurred at 600-950 hPa. The 180-hr forecast can be divided
into two periods with mostly cloudless or cloudy conditions, respectively. For the first 3-4
days (Sep. 5-8) when MCR was under cloudless or partly cloudy conditions, dust-radiation
interaction was mostly through direct effects combining both shortwave and longwave
radiation processes. In general, dust absorption of shortwave radiation creates a heating effect
around the dust maxima, and dust extinction (absorption and scattering) of shortwave
radiation reduces the downward radiative flux, thus leading to a cooling effect below the dusty
layer. For longwave radiation, dust absorbs and reflects energy from below and then re-emits
longwave radiation in all directions, yielding a cooling effect in the dusty layer and a warming
effect near the surface. It can be seen from Figure 2c that within the dusty layer the net
heating effect resulted from the dominated shortwave absorption, whereas the net warming
near the surface was caused by the dominated longwave radiation [Carlson and Benjamin,
1980; Zhu et al., 2007; Wong et al., 2009; Chen et al., 2010].
Over time, MCSs developed and the MCR was under cloudier condition from Sep. 9 as deep
and moist convection was triggered by persistent lower tropospheric convergence associated
with pre-existing disturbances. When dust interacted with clouds, the clouds significantly
influence the direct and semi-direct radiative impacts of dust [Su et al., 2008; Quijano et al,
2000] and the aerosol-induced warming due to semi-direct effect of dust is much more
pronounced than direct effect [Su et al., 2008]. This caused larger temperature increase in the
middle troposphere from 12 UTC 10 September. The increased temperature below 600 hPa
led to more stable atmosphere and suppressed the convection [Wong and Dessler , 2005;
Wong et al., 2009], as evidenced by decreased upward vertical velocity for Sep. 9-11 (Figure
2b). The dust also evaporated low-level clouds through semi-direct effect [Grassl, 1975].
However, starting from ~12 UTC 11 September, the MCR-averaged vertical velocity and
clouds were increased. We thus further divide the cloudy period into two phases: before and
after 12 UTC 11 September. For each phase, we select one time step model output to analyze
in detail the dust effects to TC development.
6 Figure 3 shows the spatial distribution of column-integrated CLWP valid at 00 UTC 10
September and 114-hr~120-hr accumulated convective precipitation in the MCR. Figure 4
displays the 10-m winds overlaid with 850-hPa potential temperature at the same time. Three
MCSs occurred in NO_DA, but MCS A centered at (20º N, 45º W) disappeared in DA. The
differences between the two experiments (right panels in Figures 4) show significant decrease
of cloud and convective rain at the location of the MCS A. That was mainly due to the dustinduced evaporative effects. As indicated in many previous studies (e.g., Leary and Houze,
1979; Brown, 1979; Sun et al., 2009), evaporative and melting cooling can create negative
buoyancy and would induce meso-scale downdrafts. Figure 5 depicts the west-east cross
sections across the center of MCS A (red dashed line in Figure 4). Figures 5a,b are the (u, 100
×vertical wind speed) vectors and potential temperature from DA and NO_DA respectively.
Figures 5c,d are the differences of potential temperature and RH between the two
experiments (DA-NO_DA). The differences of (u, 100×w) vectors (overlaid in Figure 5c)
show enhanced downward motion and a cold pool around the center of MCS A at ~44º W. ~5K potential temperature decrease and 80% relative humidity decrease occurred in the middle
troposphere where MCS A was located. The enhanced downdraft stabilized the boundary
layer and weakened the large-scale boundary layer convergence, which drives the deep and
moist convection. Over time, these processes resulted in the erosion of the MCS. The surface
wind speed around the MCS A (Figure 4) decreased by ~10 m/s. This disturbance suppression
is very similar to the WRF simulations with AIRS temperature assimilation in Sun et al.
[2009].
Figures 6-8 are the same as Figures 3-5 but for the forecast valid at 00 UTC 12 September.
Similar to the suppression of the MCS A at 00 UTC 10 September, the MCS D (centered at
16º N, 45º W) disappeared in DA with a large decrease of potential temperature and RH
(Figure 8a,b). The MCS E retained in DA but shifted a little bit northward (Figure 7).
However, the spatial distributions of CLWP and convective rainfall (Figure 6) show not only
significant decrease at the centers of the MCSs but also obvious increases in the peripheral
region (e.g., from the DA experiment, stronger rain band in the south of MCS D location).
The cross sections at MCS E also shows stronger convection in the peripheral region. Jenkins
et al. [2008] found the linkage between Saharan dust and invigoration of TC convective rain
bands by using satellite imagery and rawinsondes. The enhanced cloud water content and
cloud- and precipitation-sized particles associated with tropical cyclogenesis were observed
just south of the SAL during the NASA African Monsoon Multidisciplinary Analysis
(NAMMA) campaign (August-September 2006) in the Eastern Atlantic. Rosenfeld et al.
[2012] also used satellite observations and found that aerosols present in the peripheral clouds
of the TC slow the rain-forming processes there, invigorates the convection in peripheral
region later, and decrease the TC maximum wind speed. Consistent with the observations in
Koren et al. [2005] and Rosenfeld et al. [2012], more liquid and ice clouds in the higher layers
(around 00 UTC 12 September in Figure 2d) were invigorated in the DA experiment, because
more water can ascend to freezing levels as supercooled water due to early suppression.
7 Besides, the increased upward vertical velocity (Figure 2b) is also evident. According to the
conceptual model in Rosenfeld et al. [2012], the greater vigor of the peripheral clouds draws
more ascending air at the periphery of the storm (indicated by the strong winds between 1214º N, 35-45º W in Figure 7a), thereby bleeding the low-level airflow toward the storm center
D. In addition, the intensified ice precipitation in the peripheral clouds melted and evaporated
at the lower levels, thereby cooling the air that converges into the center (cool pool at 45º W
in Figure 8c). The cooler air has less buoyancy and hence dampens the rising air in the storm
center, thereby further weakening the convergence and the maximum wind speed of the storm.
In Sun et al. [2009], after the assimilation of AIRS temperature in the WRF simulations the
TC disturbances were also suppressed for the same event. Their analysis showed that the
major reason for disturbance suppression is the better-simulated warm and dry SAL features.
They found that the thermal structure of the SAL led to low-level temperature inversion and
increased stability and vertical wind shear, which play direct roles to the TC suppression.
They also found the warm SAL temperature may also play the indirect effects by increasing
evaporative cooling and initiating mesoscale downdrafts. Their explanations were consistent
for the disappearance of MCS A in our AOD DA experiment. The difference is that they
assimilated AIRS temperature to improve meteorological ICs, whereas we assimilated
MODIS AOD to improve the dust ICs and thus alter subsequent TC forecast via direct and
semi-direct dust-radiative feedback effects. In Sun et al. [2009], the WRF forecasts only last
for 96 hours and no dust radiative effects were considered, thus the later invigoration was not
reported.
4.3 Impact on SAL
Saharan dust, often propagating downstream along SAL to the Atlantic Ocean, can also
modify SAL by changing the energy budget through direct and semi-direct radiative effects.
In our two 180-hr forecasts initialized from the same meteorological ICs, the meteorological
differences in the two forecasts were all due to dust-radiative effects. As shown in Figure 2,
there are obvious T and RH changes when dust interacted with clouds through semi-direct
effects after Sep. 10. Figure 9 shows the spatial distributions of 850 hPa T and 700 hPa RH,
from the two experiments as well as the corresponding difference fields (DA-NO_DA) for
one-week forecast valid at 00 UTC 12 September. The tongue shape of the warm and dry SAL
extended more westward in DA with ~5 degree T increase and 70% RH decrease over the
Eastern Atlantic. Figure 12 displays the west-east cross sections of T and RH differences
along the black solid line in Figure 10.
5. Conclusions
In this study, the GSI/WRF-Chem coupled AOD DA system was used to investigate the dustradiative effects on tropical cyclogenesis and SAL over the Atlantic Ocean during a Saharan
dust outbreak event on Sep. 5-12, 2006. Prior to 180-hr forecasts initialized from 00 UTC 5
8 September, two 6-hourly cycling experiments with (DA) and without (NO_DA) assimilation
of MODIS 550-nm AOD were conducted from 00 UTC 28 August to 00 UTC 5 September.
One week AOD DA substantially improved the depiction of dust outbreak and transport, with
maximum AOD increased from ~0.25 (in NO_DA) to ~1.3 (in DA) over the Atlantic Ocean at
00 UTC 5 September. In addition, the assimilated dust-contributed AOD ratio in total AOD
was increased from < 0.4 to > 0.8. The better-analyzed dust outbreak features were consistent
with the OMI Aerosol Index images and allowed a more realistic model simulation of dust
radiative effects.
180-hr forecasts were then initialized from 00 UTC 5 September for the two experiments with
different aerosol ICs but the same meteorological ICs interpolated from the GFS analysis. In
the Main Cyclogenesis Region (MCR) over ocean during the mostly cloudless period (Sep. 57), the air was maximally heated at 600-850 hPa levels because of the dominated dustabsorption effect of shortwave solar radiation. Another local heating was found in the lower
boundary layer due to the dominance of downward longwave radiation emitted from dusty
layer. After Sep. 9, MCSs formed. Dust semi-direct effect induced much larger T increase and
RH decrease in the middle troposphere within the MCR, where the warm air enhanced the
static stability, suppressed convection and evaporated clouds during the subsequent two days
(Sep. 9-10). These changes to clouds and precipitation further resulted in T decrease and RH
increase in the lower troposphere (below 850 hPa).
In the first few days of MCS development, the warm anomaly induced by dust semi-direct
radiative effect suppressed clouds and precipitation in the high AOD region and enhanced
downward motion, and a cold pool occurred at the storm center on Sep 9-10. ~5 degrees
potential temperature decrease and 80% RH decrease occurred in the middle troposphere.
Because the same meteorological ICs were used in both DA and NO_DA experiments and
dust radiative feedback took time to be in effect, the warmness and dryness features of SAL
from the AOD DA experiment were more evident at later stages of the 180-hr forecast. An up
to 7-degree T increase and 80% RH decrease at 600-850 hPa were reached for the forecast
valid at 00 UTC 12 September. However, a coupled MET plus aerosol DA [Schwartz et al.,
2014], which simultaneously assimilates meteorological and AOD observations, is expected
to better reveal SAL features in the earlier stage of the forecast. This will be further explored
in the future with possibly more case studies in order to draw more general conclusions.
Reference
Brown, J. M. (1979), Mesoscale Unsaturated Downdrafts Driven by Rainfall Evaporation Numerical Study, J Atmos Sci, 36(2), 313-338.
Carlson, T. N., and S. G. Benjamin (1980), Radiative Heating Rates for Saharan Dust, J
Atmos Sci, 37(1), 193-213.
Chen, S. H., S. H. Wang, and M. Waylonis (2010), Modification of Saharan air layer and
environmental shear over the eastern Atlantic Ocean by dust-radiation effects, J Geophys
Res-Atmos, 115.
9 Chin, M., D. L. Savoie, B. J. Huebert, A. R. Bandy, D. C. Thornton, T. S. Bates, P. K. Quinn,
E. S. Saltzman, and W. J. De Bruyn (2000), Atmospheric sulfur cycle simulated in the
global model GOCART: Comparison with field observations and regional budgets, J
Geophys Res-Atmos, 105(D20), 24689-24712.
Chin, M., P. Ginoux, S. Kinne, O. Torres, B. N. Holben, B. N. Duncan, R. V. Martin, J. A.
Logan, A. Higurashi, and T. Nakajima (2002), Tropospheric aerosol optical thickness
from the GOCART model and comparisons with satellite and Sun photometer
measurements, J Atmos Sci, 59(3), 461-483.
Chou, M.-D., and M.J. Suarez (1994), An efficient thermal infrared radiation parameterization
for use in general circulation models, NASA Tech. Memo., TM 104606, vol. 3, 25 pp.,
NASA Goddard Space Flight Cent., Greenbelt, Md.
Collins, W. D., P. J. Rasch, B. E. Eaton, B. V. Khattatov, J. F. Lamarque, and C. S. Zender
(2001), Simulating aerosols using a chemical transport model with assimilation of
satellite aerosol retrievals: Methodology for INDOEX, J Geophys Res-Atmos, 106(D7),
7313-7336.
Dunion, J. P., and C. S. Velden (2004), The impact of the Saharan air layer on Atlantic
tropical cyclone activity, B Am Meteorol Soc, 85(3), 353-+.
Evan, A. T., J. Dunion, J. A. Foley, A. K. Heidinger, and C. S. Velden (2006), New evidence
for a relationship between Atlantic tropical cyclone activity and African dust outbreaks,
Geophys Res Lett, 33(19).
Ginoux, P., M. Chin, I. Tegen, J. M. Prospero, B. Holben, O. Dubovik, and S. J. Lin (2001),
Sources and distributions of dust aerosols simulated with the GOCART model, J
Geophys Res-Atmos, 106(D17), 20255-20273.
Gray, W. M. (1968), Global View of Origin of Tropical Disturbances and Storms, Mon
Weather Rev, 96(10), 669-&.
Grassl, H.(1975), Albedo Reduction and Radiative Heating of Clouds by Absorbing Aerosol
Particles, Contrib. Atmos. Phys., 48, 199–210
Grell, G. A., and D. Devenyi (2002), A generalized approach to parameterizing convection
combining ensemble and data assimilation techniques, Geophys Res Lett, 29(14).
Grell, G. A., S. E. Peckham, R. Schmitz, S. A. McKeen, G. Frost, W. C. Skamarock, and B.
Eder (2005), Fully coupled "online" chemistry within the WRF model, Atmos Environ,
39(37), 6957-6975.
Han, Y., P. van Delst, Q. Liu, F. Weng, B. Yan, R. Treadon, and J. Derber (2006), JCSDA
Community Radiative Transfer Model (CRTM)- Version 1, NOAA Tech. Rep. NESDIS
122, 33 pp., NOAA Silver Spring, Md.
Hsu, N. C., S. C. Tsay, M. D. King, and J. R. Herman (2004), Aerosol properties over brightreflecting source regions, Ieee T Geosci Remote, 42(3), 557-569.
Hsu, N. C., S. C. Tsay, M. D. King, and J. R. Herman (2004), Aerosol properties over brightreflecting source regions, Ieee T Geosci Remote, 42(3), 557-569.
Jenkins, G. S., A. S. Pratt, and A. Heymsfield (2008), Possible linkages between Saharan dust
and tropical cyclone rain band invigoration in the eastern Atlantic during NAMMA-06,
Geophys Res Lett, 35(8).
Jones, T. A., D. J. Cecil, and J. Dunion (2007), The environmental and inner-core conditions
governing the intensity of Hurricane Erin (2001), Weather Forecast, 22(4), 708-725.
10 Karyampudi, V. M., and T. N. Carlson (1988), Analysis and Numerical Simulations of the
Saharan Air Layer and Its Effect on Easterly Wave Disturbances, J Atmos Sci, 45(21),
3102-3136.
Karyampudi, V. M., and H. F. Pierce (2002), Synoptic-scale influence of the Saharan air layer
on tropical cyclogenesis over the eastern Atlantic, Mon Weather Rev, 130(12), 31003128.
Koren, I., Y. J. Kaufman, D. Rosenfeld, L. A. Remer, and Y. Rudich (2005), Aerosol
invigoration and restructuring of Atlantic convective clouds, Geophys Res Lett, 32(14).
Landsea, C. W. (1993), A Climatology of Intense (or Major) Atlantic Hurricanes, Mon
Weather Rev, 121(6), 1703-1713.
Lau, K. M., K. M. Kim, Y. C. Sud, and G. K. Walker (2009), A GCM study of the response
of the atmospheric water cycle of West Africa and the Atlantic to Saharan dust radiative
forcing, Ann Geophys-Germany, 27(10), 4023-4037.
Leary, C. A., and R. A. Houze (1979), Melting and Evaporation of Hydrometeors in
Precipitation from the Anvil Clouds of Deep Tropical Convection, J Atmos Sci, 36(4),
669-679.
Liu, Q. H., and F. Z. Weng (2006), Advanced doubling-adding method for radiative transfer
in planetary atmospheres, J Atmos Sci, 63(12), 3459-3465.
Liu, Z. Q., Q. H. Liu, H. C. Lin, C. S. Schwartz, Y. H. Lee, and T. J. Wang (2011), Threedimensional variational assimilation of MODIS aerosol optical depth: Implementation
and application to a dust storm over East Asia, J Geophys Res-Atmos, 116.
Parrish, D. F., and J. C. Derber (1992), The National-Meteorological-Centers Spectral
Statistical-Interpolation Analysis System, Mon Weather Rev, 120(8), 1747-1763.
Pratt, A. S., and J. L. Evans (2009), Potential Impacts of the Saharan Air Layer on Numerical
Model Forecasts of North Atlantic Tropical Cyclogenesis, Weather Forecast, 24(2), 420435.
Quijano, A. L., I. N. Sokolik, and O. B. Toon (2000), Radiative heating rates and direct
radiative forcing by mineral dust in cloudy atmospheric conditions, J Geophys ResAtmos, 105(D10), 12207-12219.
Ramanathan, V., P. J. Crutzen, J. T. Kiehl, and D. Rosenfeld (2001), Atmosphere - Aerosols,
climate, and the hydrological cycle, Science, 294(5549), 2119-2124.
Reale, O., W. K. Lau, K. M. Kim, and E. Brin (2009), Atlantic Tropical Cyclogenetic
Processes during SOP-3 NAMMA in the GEOS-5 Global Data Assimilation and Forecast
System, J Atmos Sci, 66(12), 3563-3578.
Remer, L. A., et al. (2005), The MODIS aerosol algorithm, products, and validation, J Atmos
Sci, 62(4), 947-973.
Remer, L. A., et al. (2005), The MODIS aerosol algorithm, products, and validation, J Atmos
Sci, 62(4), 947-973.
Rosenfeld, D., W. L. Woodley, A. Khain, W. R. Cotton, G. Carrio, I. Ginis, and J. H. Golden
(2012), Aerosol Effects on Microstructure and Intensity of Tropical Cyclones, B Am
Meteorol Soc, 93(7), 987-1001.
Schwartz, C. S., Z. Q. Liu, H. C. Lin, and S. A. McKeen (2012), Simultaneous threedimensional variational assimilation of surface fine particulate matter and MODIS
aerosol optical depth, J Geophys Res-Atmos, 117.
Shu, S. J., and L. G. Wu (2009), Analysis of the influence of Saharan air layer on tropical
cyclone intensity using AIRS/Aqua data, Geophys Res Lett, 36.
11 Sokolik, I. N., O. B. Toon, and R. W. Bergstrom (1998), Modeling the radiative
characteristics of airborne mineral aerosols at infrared wavelengths, J Geophys ResAtmos, 103(D8), 8813-8826.
Su, J., J. P. Huang, Q. Fu, P. Minnis, J. M. Ge, and J. R. Bi (2008), Estimation of Asian dust
aerosol effect on cloud radiation forcing using Fu-Liou radiative model and CERES
measurements, Atmos Chem Phys, 8(10), 2763-2771.
Su, J., J. P. Huang, Q. Fu, P. Minnis, J. M. Ge, and J. R. Bi (2008), Estimation of Asian dust
aerosol effect on cloud radiation forcing using Fu-Liou radiative model and CERES
measurements, Atmos Chem Phys, 8(10), 2763-2771.
Sun, D. L., K. M. Lau, and M. Kafatos (2008), Contrasting the 2007 and 2005 hurricane
seasons: Evidence of possible impacts of Saharan dry air and dust on tropical cyclone
activity in the Atlantic basin, Geophys Res Lett, 35(15).
Sun, D. L., W. K. M. Lau, M. Kafatos, Z. Boybeyi, G. Leptoukh, C. W. Yang, and R. X.
Yang (2009), Numerical Simulations of the Impacts of the Saharan Air Layer on Atlantic
Tropical Cyclone Development, J Climate, 22(23), 6230-6250.
Wong, S., and A. E. Dessler (2005), Suppression of deep convection over the tropical North
Atlantic by the Saharan Air Layer, Geophys Res Lett, 32(9).
Wong, S., A. E. Dessler, N. M. Mahowald, P. Yang, and Q. Feng (2009), Maintenance of
Lower Tropospheric Temperature Inversion in the Saharan Air Layer by Dust and Dry
Anomaly, J Climate, 22(19), 5149-5162.
Wu, W. S., R. J. Purser, and D. F. Parrish (2002), Three-dimensional variational analysis with
spatially inhomogeneous covariances, Mon Weather Rev, 130(12), 2905-2916.
Wu, W. S., R. J. Purser, and D. F. Parrish (2002), Three-dimensional variational analysis with
spatially inhomogeneous covariances, Mon Weather Rev, 130(12), 2905-2916.
Wu, L. G. (2007), Impact of Saharan air layer on hurricane peak intensity, Geophys Res Lett,
34(9).
Zhao, C., X. Liu, L. R. Leung, B. Johnson, S. A. McFarlane, W. I. Gustafson, J. D. Fast, and
R. Easter (2010), The spatial distribution of mineral dust and its shortwave radiative
forcing over North Africa: modeling sensitivities to dust emissions and aerosol size
treatments, Atmos Chem Phys, 10(18), 8821-8838.
Zhu, A., V. Ramanathan, F. Li, and D. Kim (2007), Dust plumes over the Pacific, Indian, and
Atlantic oceans: Climatology and radiative impact, J Geophys Res-Atmos, 112(D16).
Zipser, E. J., et al. (2009), The Saharan Air Layer and the Fate of African Easterly Waves, B
Am Meteorol Soc, 90(8), 1137-1156.
12 Figure 1. After one-week period (28 August – 5 September 2006) for aerosol spin-up, total
550-nm AOD distribution valid at 00 UTC 5 Sep. 2006 from the (a) NO_DA and (b) DA
experiments. Also shown are dust AOD ratios in red contours. Note that color scales are
intentionally made different between NO_DA and DA for a better visualization. Black
rectangle in (b) is defined as the Main Cyclogenesis Region (MCR).
Figure 2. MCR-averaged differences (DA - NO_DA) for (a) AOD, (b) vertical velocity, (c)
temperature (color shaded) and relaive humidity (contours), and (d) liquid and ice clouds as a
function of height and time.
13 Figure 3. Spatial distribution and differences of (a) column integrated clouds liquid water path
and (b) 114-~120-hr convective rainfall in MCR from two experiments for 120-hr forecast
valid at 00UTC 10 Sep. Units: mm.
Figure 4. 10-m u-v wind vectors (m/s) and potential temperatures (shaded, K) in MCR for
120-h forecast valid at 00 UTC 10 Sep. The red dashed lines are the west-east cross-section
used in Fig. 5.
14 Figure 5. Cross sections (along red dashed line in Fig. 6) of (u, 100×W) vectors and potential
temperature from the (a) DA and (b) NO_DA experiments as well as (c) the corresponding
differences (DA – NO_DA) and (d) relative humidity difference of DA minus NO_DA.
15 Figure 6. Same as Fig. 3 but for 168-hr forecast valid at 00 UTC 12 Sep.
Figure 7. Same as Fig. 4 but for 168-hr forecast valid at 00 UTC 12 Sep. The red dashed lines
are the west-east cross section used in Fig. 8.
16 Figure 8. Same as Fig. 5 but for 168-hr forecast valid at 00 UTC 12 Sep. The cross sections
are defined as the red dashed lines in Fig. 7.
17 Figure 9. Spatial distributions of (a) 850 hPa temperature (K) and (b) 750 hPa relative
humidity (%) from two experiments for 168-hr forecast valid at 00 UTC 12 Sep and the their
corresponding differences. The black solid line in the difference field is for the cross sections
used in Figure 10.
Figure 10. Cross sections (along black solid line in Figure 9) of (a) temperature difference and
(b) relative humidity differences of DA minus NO_DA for 168-hr forecast valid at 00 UTC 12
Sep. 2006.
18 
Download